You are on page 1of 18

Journal of Sedimentary Research, 2011, v.

81, 248–265
Perspectives
DOI: 10.2110/jsr.2011.24

DOLOMITE IN PERMIAN PALEOSOLS OF THE BRAVO DOME CO2 FIELD, U.S.A.: PERMIAN REFLUX
FOLLOWED BY LATE RECRYSTALLIZATION AT ELEVATED TEMPERATURE

KATHERINE A. HARTIG,1 GERILYN S. SOREGHAN,2 ROBERT H. GOLDSTEIN,3 AND MICHAEL H. ENGEL2


1
Pioneer Natural Resources, 5205 North O’Connor Boulevard, Irving, Texas 75039, U.S.A.
e-mail: Katherine.Hartig@pxd.com
2
Conoco-Phillips School of Geology and Geophysics, University of Oklahoma, 100 East Boyd Street, Norman, Oklahoma 73019, U.S.A.
3
Department of Geology, University of Kansas, 1475 Jayhawk Boulevard, Lawrence, Kansas 66045, U.S.A.

ABSTRACT: Lower Permian Abo–Tubb subsurface strata of northeastern New Mexico, U.S.A. consist of 30 to 120 m of fluvial
and eolian deposits, containing numerous dolomitic paleosols. Dolomite in the Abo–Tubb interval occurs in discrete intervals,
interpreted as paleosols on the basis of recognizable horizonation and the presence of associated features such as ped structures
and rhizoliths. The dolomite commonly forms veins, rhizoliths, and discrete and coalesced nodules.
Petrographic analysis indicates four dolomite types present in paleosol horizons: microsparitic, pseudospherulitic, clear cement,
and cloudy (inclusion-rich) cement. Microsparitic dolomite constitutes most of (70%) the dolomite, and is stoichiometric, with
relatively low Sr and Na and elevated Fe and Mn. Pseudospherulitic dolomite forms 15% of the dolomite, and consists of
stoichiometric microspar and spar, with abundant fluid inclusions, low Sr, Na, and Fe, and elevated Mn. Clear dolomite cement
constitutes approximately 12% of the dolomite and is finely to coarsely crystalline with a relatively calcian composition, low Sr and
Na, and elevated Fe and Mn. The cloudy cement forms a minor (2%) phase of very coarsely crystalline rhombs with a calcian
composition, low Sr and Na values, and relatively high Fe and Mn. All dolomite types exhibit a mottled texture of moderate to
bright orange and red luminescence. Fluid-inclusion analysis was not possible on the most common (microsparitic) dolomite type,
but analyses on the coarser varieties reveal homogenization temperatures typically ranging from 70uC to 110uC and final melting
temperatures of ice ranging from 220uC to 234uC, suggesting fluids of high (25% by weight) salinity. Bulk-rock carbon and
oxygen isotopes of carbonate nodules and rhizoconcretions from four paleosol profiles most commonly exhibit mean d13C and d18O
values typically ranging between , 23 to 25.5% (VPDB) and 20.5 to 2.8% (VPDB), respectively. Sulfur isotopes were analyzed
for the overlying Cimarron Anhydrite interval and an anhydrite nodule near the base of the study interval. Both samples yielded
similar results, suggesting that brines that precipitated the Cimarron Anhydrite also influenced the study interval.
The dolomite in the Abo–Tubb interval is interpreted to have precipitated in multiple stages. Solid calcite microinclusions in
microsparitic and pseudospherulitic dolomite suggest that pedogenensis and groundwater penecontemporaneous with Early
Permian deposition led to precipitation of calcite rather than dolomite. Oxygen and sulfur isotope data indicate dolomitization,
from later reflux of brines associated with deposition of the overlying (Leonardian) Cimarron Anhydrite. Petrographic relations
and elevated temperatures and salinities recorded by the fluid inclusions in the coarse dolomite are consistent with
recrystallization of the earlier dolomite and precipitation of dolomite cement from warm saline fluids of a possible Tertiary
hydrothermal origin, just predating migration of mantle-derived gas into the Bravo Dome reservoir.
These findings offer insight into questions about the climate significance of dolomite in paleosols, origin of dolomite in
hydrothermally altered hydrocarbon reservoirs, and the long-term effect of CO2 sequestration in saline aquifers. The results
show that dolomitic nodules in paleosols may form as replacements of paleosol calcite long after the paleosol formed. Primary
pedogenic dolomite would undoubtedly have climatic and paleoenvironmental significance, but recognizing it as forming during
soil formation, as opposed to later dolomitization, is key in establishing that significance. Also, this study provides an example
of high-temperature recrystallization of low-temperature dolomite, a process hotly debated in discussions of dolomitized
hydrocarbon reservoirs with evidence for hydrothermal fluid flow. Finally, as a natural CO2 reservoir, the Bravo Dome system
shows that the long-term result of dissolution of CO2 into brines is predominantly carbonate mineral dissolution.

INTRODUCTION Lith et al. 2003; Moreira et al. 2004; Roberts et al. 2004: Kenward et al.
2009). Nevertheless, precipitation of primary dolomite has been
Dolomite is more abundant in ancient carbonates than modern documented in marine settings (sabkhas, lagoonal, and open), saline
carbonate environments, and laboratory synthesis of dolomite at surface evaporative lakes, and other low-temperature sedimentary environments
conditions is commonly unsuccessful (Hardie 1987) unless there is a (McKenzie et al. 1980; Garces and Aguilar 1994; Wanas 2002). In these
microbial input overcoming kinetic barriers (Vasconcelos et al. 1995; Van settings, dolomite precipitates from seawater or lake water with high

Copyright E 2011, SEPM (Society for Sedimentary Geology) 1527-1404/11/081-248/$03.00


JSR PERSPECTIVES 249

Mg2+ concentrations. Additionally, primary, biogenically mediated and enhanced oil recovery continue as subjects of active research
dolomite precipitation has been documented in anoxic, isolated lagoons (Broadhead 1990, 1998). The Bravo Dome field is long-term analog for
(Vasconcelos et al. 1995) and in evaporative lakes as a result of bacterial geologic CO2 sequestration. It results from geologically recent CO2
sulfate reduction (Wright and Wacey 2005), with laboratory replication of injection and thus provides ground truth for modeling efforts. Previous
both processes (Vasconcelos et al. 1995; Wright 1999; Wright and Wacey research has shown that long-term analog studies are essential for
2005; Wacey et al. 2007). calibration of reactive transport models in such systems (e.g., Pearce et al.
The occurrence of primary dolomite in modern soils has been 1996; Baines and Worden 2004). In particular, this study adds essential
documented only rarely, and typically in association with Mg-rich parent analog data on the reactivity of mineral phases in an arkosic aquifer used
material (e.g., basaltic host; Capo et al. 2000; Whipkey et al. 2002; for geologic CO2 sequestration.
Whipkey and Hayob 2008) or induced by highly evaporative meteoric
and groundwater soil solutions (Kohut et al. 1995). In contrast, much of GEOLOGIC SETTING
the dolomite present in paleosols (ancient soils) has been attributed to
groundwater processes. Specifically, elevated evaporation rates coupled The Lower Permian Abo Formation and the overlying Yeso Formation
with various groundwater processes have been documented to produce occur in the shallow subsurface of northeastern New Mexico adjacent to the
Mg-rich groundwater capable of either precipitating dolomite or Bravo Dome paleohigh. The Bravo Dome uplift represents the southeast
dolomitizing a precursor calcite. Specific processes resulting in Mg-rich plunging projection of the Sierra Grande uplift, which developed as a part
groundwater include: (1) mixing of saline brines and fresh groundwater of the extensive late Paleozoic deformation related to the greater Ancestral
(Searl 1988; El-Sayed et al.1991; Coloson and Cojan 1996); (2) calcrete Rocky Mountains system (Fig. 1; Roberts et al. 1976; Ross and Ross 1986;
precipitation from nearby groundwater resulting in groundwater depleted Broadhead and King 1988; Broadhead 1990). The Dalhart, Palo Duro, and
in Ca2+ and enriched in Mg2+ (Botha and Hughes 1992; Armenteros et al. Tucumcari basins formed adjacent to Bravo Dome. The Bravo Dome
1995; Spötl and Wright 1992); (3) fluid movement through Mg-rich clays merges northwest with the Sierra Grande uplift and southeast with the
(Pimentel 2002); and (4) introduction of brackish waters during post- Amarillo uplift (Fig. 1). Paleogeographic reconstructions place the Bravo
exposure marine transgressions in paleosols hosted on marine carbonate Dome region in low, equatorial latitudes (, 0–10uS to N) during
(Wright 1999; Wright et al. 1997; Barnett et al. in press). As many Pennsylvanian and Early Permian time (Scotese 1999; Kessler et al. 2001).
researchers have interpreted dolomite in paleosols as an indicator of The Abo–Tubb study interval consists of the Abo Formation and the
climate or atmospheric conditions (e.g., Wright and Tucker 1991; Spötl overlying Tubb Sandstone Member of the Yeso Formation, hereafter
and Wright 1992; Kessler et al. 2001; Alonso-Zarza 2003; Benito et al. referred to as the ‘‘Abo–Tubb’’ interval. In the Bravo Dome region, the
2005), it is necessary to determine if dolomite in paleosols formed during Abo–Tubb interval rests nonconformably on Precambrian basement and
pedogenesis, or as a result of later diagenetic alteration. is capped by the Cimarron Anhydrite Member of the Yeso Formation
Diagenetic processes, both early and late, have been shown to (Fig. 2). The Abo–Tubb interval consists of orange and red to purple
contribute to dolomite formation in paleosols. Specific mechanisms siliciclastic strata of fluvial and eolian origin, with numerous interbedded
associated with late diagenetic events might include: (1) movement of dolomitic paleosols (Kessler et al. 2001). Interval thicknesses throughout
warm hydrothermal fluids capable of precipitating and recrystallizing the study area vary from 30 to 120 m, reflecting the subjacent
dolomite (Gregg 1985; Gregg and Shelton 1990); and (2) burial processes paleotopography of Bravo Dome (Fig. 3; Kessler et al. 2001). The age
lacking such fluid movement (Gao and Land 1991; Montañez and Read of the Abo–Tubb interval is early Wolfcampian to early Leonardian (see
1992; Smith and Dorobek 1993). Kessler et al. 2001 for details).
This paper addresses the origin of dolomite in paleosols of lower The Abo–Tubb interval in the Bravo Dome forms the reservoir in a 10
Permian fluvial and eolian deposits of northeastern New Mexico, U.S.A. TCF CO2 field (Broadhead 1990, 1993). The gas had a mantle source
Much of the dolomite present in this system exhibits spatial and textural (Staudacher 1987; Cassidy 2005; Gilfillan et al. 2008) and was likely
attributes consistent with a pedogenic origin. Consideration of a variety derived from degassing of basaltic magmas dating from 1 Ma to 8 ka,
of petrographic and geochemical attributes together with stratigraphic with CO2 charging possibly continuing to the present (Broadhead 1998;
associations, however, suggest dolomitization of precursor calcite during Baines and Worden 2004; Broadhead and Gillard 2004; Cassidy 2005).
an early-stage seepage of evaporated (marine) brines, followed by later
diagenetic events of recrystallization and cementation attributed to Depositional and Pedogenic Facies of the Abo–Tubb Interval
tectonic and hydrothermal processes. This work provides a new model of
dolomite in paleosols and thus broadens documentation and understand- This paper focuses specifically on the dolomite in paleosols of the Abo–
ing of the origin of dolomite in paleosols. Tubb interval and builds on the detailed depositional and paleopedolo-
Through detailed petrographic and geochemical study, this work sheds gical framework established by Kessler et al. (2001; outlined below).
light on the origin of some hydrothermal dolomite. Hydrothermal Kessler et al. (2001) documented three primary depositional facies in the
dolomitization is a popular model for explaining massively dolomitized study interval: (1) matrix-supported conglomerate deposits inferred to be
oil and gas reservoirs (e.g., Smith and Davies 2006; Davies and Smith of fluvial origin and typically found near the base of the section; (2)
2006) and large volumes of ancient dolomite globally. Concerns about planar-bedded to cross-bedded sandstone deposits inferred to be of fluvial
sources of Mg and the nature of geochemical data have led to suggestions origin, predominating in the middle of the interval; and (3) massive to
of high-temperature recrystallization of low-temperature precursor faintly laminated siltstone loess deposits predominant in the top of the
dolomite in such systems (Machel 2004). This study provides evidence section. Thin (, 30 cm) intraclastic layers consisting of reworked mud
of recrystallization of low-temperature dolomite at elevated temperatures. and pedogenic nodule rip-ups occur in fluvial strata.
Finally, this study elucidates fluid flow and thermal history of one of On average, paleosols occur at 5 m intervals in the thickest (off-
the most important CO2 reservoirs of North America, the Bravo Dome of structure) Abo–Tubb intervals and are more closely spaced in the thinnest
New Mexico. It adds important constraints on the tectonic and thermal (on-structure) areas (Figs. 3, 4, 5). Paleosols range from 10 cm to 4 m
history, to allow the gas migration to be placed into the context of a more thick and are recognized by the presence of downwardly bifurcating
extensive paragenesis, and thus tests hypotheses of long-term effects of (inferred root) traces, ped structures, slickensides (on ped surfaces),
geologic CO2 sequestration in the subsurface. Both the geologic melanization (darkening), and carbonate nodules and rhizoconcretions.
sequestration of CO2, and CO2 as a resource for industrial purposes Kessler et al. (2001) described three paleosol types: Vertic Dolosol, Vertic
250 K.A. HARTIG ET AL. JSR

FIG. 1.—Major late Paleozoic tectonic fea-


tures in the greater study area, southwestern
U.S.A. Patterned areas represent uplifts. Boxed
region around Bravo Dome denotes the study
area. Diagonal line represents the paleolatitude
line from Scotese (1999) (modified from Kessler
et al. 2001).

Protosol, and Dolic Protosol (terms slightly modified from the Mack et
al. 1993 classification scheme). The prefixes ‘‘dolo and dolic’’ are utilized
to highlight the presence of pedogenic carbonate textures in the paleosol,
which currently consist of dolomite. The prefix ‘‘vertic’’ is used to refer to
the presence of pedogenic slickensides. Vertic Dolosols typically
predominate in the lower one-third of the study interval, associated with
the fluvial conglomeratic facies, and they range in thicknesses from 20 cm
to 4 m. Vertic Protosols occur in the upper half of the study interval,
associated with the fluvial sandstone and eolian siltstone facies, and range
up to 1.2 m thick. Dolic Protosols typically occur in the upper half of the
study interval, range from 10 cm to 2 m thick, and are developed in
association with the eolian siltstone (loessitic) facies.

METHODS

The study area contains a natural CO2 reservoir with abundant wells
and associated conventional cores. For this study, four wells were selected
from the thicker Abo–Tubb strata in the study area, since thicker
intervals are less likely to contain superimposed paleosols. The four
selected cores (Pittard #1, Nikkel #1, State F.T., and Libby #1; Figs. 4,
5) were logged, sampled, and photographed.
All macroscopic dolomite types were sampled from the selected cores at
various depths for petrographic analysis. Thirty polished thin sections
were stained with alizarin red-S and analyzed using both transmitted light
and cathodoluminescence (CL) to document microscopic attributes such
as crystal size, texture, and fabrics. Luminescence attributes were
obtained using a Technosyn cathodoluminescence microscope (with 9
to 11 kV and 80 to100 mA).
A subset of eight thin sections selected from Vertic Dolosol, Dolic
FIG. 2.—Permian stratigraphy of the study region from Broadhead and King Protosol, and intraclastic zones at various depths of the Pittard #1 and
(1988) and Kessler et al. (2001). Libby #1 cores, and containing all recognizable dolomite forms, were
JSR PERSPECTIVES 251

FIG. 3.—Fence diagram through studied wells


showing correlation of cored interval and general
depositional facies. See Figure 4 for well loca-
tions (modified from Kessler et al. 2001).

selected for fluid-inclusion analysis (Fig. 5). Doubly polished thick (60 mm al. 1990) and the software of Bakker (2003). Because detailed ionic ratios
thick) sections were impregnated with blue resin and mounted using cold- were not determined in this study, a model composition was assumed for
curing epoxy. The fluid-inclusion sections were cut into small chips to calculation of salinities (e.g., Goldstein and Reynolds 1994), using the
isolate targeted inclusions and placed in acetone to remove the epoxy. A mean Na:Ca ratio analyzed from 82 oil-field brines (Carpenter et al. 1974;
Linkam fluid-inclusion stage was utilized to measure homogenization Na:Ca wt. ratio of 2.1). Dissolved sulfate was not considered in
temperatures and low-temperature phase changes, using cycling methods calculating salinity. Assumptions and calibrations for low-temperature
described in Goldstein and Reynolds (1994). Bulk salinities of fluid phase equilibria lead to minor errors in determination of fluid salinity.
inclusions were calculated from measured final melting temperatures Approximately 72 inclusions were analyzed, resulting in 72 homogeniza-
using published calibrations for the NaCl–CaCl2–H2O system (Oakes et tion temperatures, 47 final melting temperatures of ice (Tm ice), and 4

FIG. 4.—Location map of wells in the Bravo


Dome study region in northeastern New Mexico,
with studied wells highlighted (modified from
Kessler et al. 2001).
252 K.A. HARTIG ET AL. JSR

FIG. 5.—Graphic columns of the four selected cores highlighting sample locations. Datum horizon is the base of the Cimarron Anhydrite.
JSR PERSPECTIVES 253

FIG. 6.—Photographs of macroscopic dolomite types. A) Dolomite rhizocretion and spar-filled rhizomolds in mottled red siltstone matrix (fine divisions are mm), B)
Dolomite calcrete nodules showing autoclastic fabrics typical of paleosols, in red siltstone matrix (fine divisions are mm), C) Dolomite, calcrete-filled fractures and
rhizoliths (scale in cm).

eutectics. All data were collected with careful constraint on their isotopes to discriminate the details of each diagenetic phase, because the
petrographic association, using the fluid-inclusion assemblage (FIA) close spatial association of phases precluded the latter. Powdered samples
approach of Goldstein and Reynolds (1994) and Goldstein (2003). The were digested in 100% phosphoric acid in sealed tubes under vacuum for
FIA approach evaluates fluid-inclusion data on the basis of the most 90 minutes at 100uC. The resultant CO2 gas was cryogenically isolated
finely discriminated events of fluid-inclusion entrapment that can be and then analyzed with a Finnigan Delta E isotope-ratio mass
identified petrographically. spectrometer. Values are reported in per mil relative to PDB carbonate.
Microprobe analysis was conducted on several thin sections selected from The d18O values are corrected for fractionation at the 100uC run
the Vertic Dolosol, Dolic Protosol, and intraclastic zones and containing all conditions (Rosenbaum and Sheppard 1986).
dolomite types. Thin sections were polished and carbon coated prior to Sulfur isotope analysis was conducted on two anhydrite samples. The
analysis. Samples were characterized at the University of Oklahoma first sample was taken from the base of the overlying Cimarron Anhydrite
utilizing a Cameca SX50 electron probe microanalyzer, equipped with a in the Nikkel #1 well. The second was sampled from an anhydrite nodule
Kevex Delta-class energy-dispersive X-ray analyzer. Phase identification located near the base of the study interval in the Libby #1 well. Samples
and textural relations were examined by backscattered electron imaging were powdered using a mortar and pestle and sent to Coastal Science
coupled with energy-dispersive X-ray analysis. Quantitative analysis was Laboratories for analysis.
performed by wavelength-dispersive spectrometry using beam conditions of
20 kV acceleration, 10 nA sample current, and a defocused 5 mm spot. DOLOMITE ATTRIBUTES
Natural crystalline solids were used as standard materials, counting times
Macroscopic Attributes
were 45 s on peak for all elements, and matrix reduction used the PAP
method of Pouchou and Pichoir (1985). Calculated elemental thresholds are Dolomite varies in concentration and morphology throughout all
as follows: 180 ppm (Ca), 245 ppm (Mg), 385 ppm (Sr), 133 ppm (Na), paleosol types in the Abo–Tubb interval. Dolomite is most abundant in
245 ppm (Fe), 385 ppm (Mn), and 110 ppm (Si). the lower half of the study interval, and it typically occurs in nodules,
Four complete Vertic Dolosol profiles (Fig. 5) were logged at rhizoconcretions, and pores.
centimeter-scale resolution and sampled for stable-isotope (C, O) Dolomite nodules are spherical to tuberose, exhibit sharp to diffuse
analyses. Core plugs (2 mm diameter) were drilled at 5 to 10 cm intervals, boundaries, and are most common in Vertic Dolosols and Dolic
targeting dolomite nodules and rhizoconcretions. Cores plugs were placed Protosols and are rare in Vertic Protosols. Nodules range in size from
in the hood and allowed to air-dry overnight. Samples were then , 1 cm to 10 cm in diameter, with the largest nodules occurring near the
individually powdered using a mortar and pestle. These bulk samples are base and the smaller discrete nodules and rhizoliths present in the upper
meant to homogenize the material present rather than to use the stable and middle parts of the study sections (Fig. 6). Nodules typically occur as
254 K.A. HARTIG ET AL. JSR
JSR PERSPECTIVES 255

TABLE 1.— Summary of geochemical data.

Mg % Na ppm
Dolomite Types Number of Samples Mean Range Std. Dev. (n) Mean Range Std. Dev.
Microsparitic 69 49 47.6 to 52.4 0.7 52 243 BDL to 988 130
Pseudospherulitic 10 48.5 4802 to 49.0 0.3 4 237 BDL to 368 92
Clear coarse cement 20 43.2 36.1 to 49.4 6 15 460 BDL to 965 316
Coarse cloudy 15 45.6 39.2 to 49.3 3.8 8 658 BDL to 2033 582
Mn ppm Fe ppm
Dolomite Types Number of Samples Mean Range Std. Dev. (n) Mean Range Std. Dev.
Microsparitic 69 1675 802 to 6504 691 55 1127 BDL to 6150 1006
Pseudospherultic 10 2733 2433 to 3382 265 2 * BDL to 6722 *
Clear coarse cement 20 2838 1618 to 4526 935 18 21824 BDL to 50498 19627
Coarse cloudy 15 5206 1964 to 8505 1883 7 41433 BDL to 70442 20761
* Indicates that statistics were not run due to the majority of the data values being below detection level (BDL).

vertically oriented cylindrical concretions and coalesced nodules (Gile et recrystallization. Fluid-inclusion analysis was not conducted on micro-
al. 1966; Bachmann and Machette 1977; Machette 1985; Kessler et al. sparitic dolomite owing to the small inclusion size (, 2 mm).
2001). Vertically aligned nodules, which are , 1 cm in diameter, likely
represent examples of carbonate concretions around anastomosing root Pseudospherulitic Dolomite.—Pseudospherulitic dolomite is most com-
traces (rhizoconcretions; Gile et al. 1966; Bachmann and Machette 1977; mon in the lower half of the study interval (adjacent to the Precambrian
Machette 1985; Kessler et al. 2001). Dolomite rhizoconcretions are basement contact). Pseudospherulitic dolomite constitutes approximately
typically , 1 cm in width, and are abundant in the upper layers of 15% of the dolomite in the study interval, and typically it is found in
paleosol profiles (Bk horizons), with relative abundance diminishing with nodules and locally in rhizoconcretions. Crystals are polymodal, range
depth. from 40 mm (finely crystalline) to 300 mm (coarsely crystalline), and
consist of nonplanar microspar and spar with anhedral to subhedral
Petrographic Attributes dolomite rhombs in a xenotopic mosaic. This dolomite is typically tan to
clear in transmitted light, and it contains abundant fluid inclusions (2 mm
Using petrographic analysis, the following four distinct dolomite types to 10 mm; Fig. 7C) that appear aligned in a radiating pattern originating
were documented on the basis of crystal size, clarity, morphology, from the center of the crystal, thus producing a pseudospherulitic texture.
presence of fluid inclusions, and cathodoluminescence attributes: (1) Cathodoluminescence reveals an overall mottled texture consisting of dull
microsparitic dolomite, (2) pseudospherulitic dolomite, (3) coarse cloudy to moderate orange and red luminescence. Areas with the brightest
dolomite, and (4) dolomite cement. luminescence are fluid-inclusion-rich zones, crystal rims, and cracks along
cleavage planes (Fig. 7D). Microprobe analysis indicates that pseudo-
Microsparitic Dolomite.—Microsparitic dolomite is the most common spherulitic dolomite is relatively stoichiometric (48.5% Mg and 51.5% Ca;
type of dolomite. It occurs throughout the study interval in rhizoconcre- Table 1) with low Sr (BDL), Na (mean 5 237) and Fe (BDL), and
tions and nodules and constitutes , 70% of all dolomite. Crystals range elevated Mn (mean 5 2733; Table 1).
in size from , 2.0 mm (cryptocrystalline) to 100 mm (medium crystalline) Two types of inclusions occur in the pseudospherulitic dolomite: (1)
and are typically equant, planar, subhedral rhombs in an idiotopic or two-phase (liquid and gas) fluid inclusions (Fig. 8); and (2) solid calcite
hypidiotopic mosaic. Crystals are tan to gray in transmitted light and inclusions. The radiating pattern and textures of the fluid inclusion
contain a moderate number of small (, 2 mm; Fig. 7A) fluid inclusions. cavities suggest a microbial origin similar to that of Rossi and Cañaveras
Cathodoluminescence analysis reveals moderate to bright orange and red (1999). Their distribution and orientation indicate a primary origin.
luminescence compared to coarser crystals, with areas of the brightest Typically, these fluid inclusions appear in two dimensions with an area of
luminescence present along crystal boundaries (Fig. 7B). approximately 20% gas and 80% liquid. Fluid-inclusion analysis was
Microprobe analysis indicates that microsparitic dolomite is approx- conducted on two fluid-inclusion assemblages (FIA), resulting in eight
imately stoichiometric (49% Mg, 51% Ca: Table 1) with relatively low homogenization temperatures (Th). The narrow range of Th values
amounts of Sr (BDL), Na (243 ppm), and Fe (1127 ppm). Manganese indicates consistent temperatures in FIAs (Goldstein and Reynolds 1994)
(Mn; 1675 ppm) values are relatively high but lower than Mn values from 77uC to 91uC with a mean value of 84.1uC (Table 2). Where FIAs
measured from the other three types of dolomite. yield consistent Th, it is likely that fluid inclusions have not been altered
There are two types of inclusions in the microsparitic dolomite: (1) two- by thermal reequilibration (Goldstein and Reynolds 1994). Eight ice final
phase (liquid and gas) fluid inclusions, and (2) solid calcite inclusions. The melting temperatures were determined and ranged from 220uC to 231uC,
distribution of fluid inclusions and their orientation suggest they are with a mean of 226.3uC; one possible eutectic was measured at 243uC.
primary, either from initial precipitation of the dolomite or its The low eutectic supports the assumed model composition of an NaCl–

FIG. 7.—Photomicrographs of dolomite types. A) Microsparitic dolomite in transmitted light. B) Microsparitic dolomite in CL. C) Pseudospherulitic dolomite in plane
light. D) Pseudospherulitic dolomite in CL. E) Dolomite cement in plane light; black line in pore represents a bubble in the epoxy. F) Dolomite cement in CL. G) Coarse
cloudy dolomite in plane light. H) Coarse cloudy dolomite in CL.
256
K.A. HARTIG ET AL.

FIG. 8.—Photomicrographs of fluid inclusions in three dolomite types. A1) Microsparitic dolomite, A2) microsparitic dolomite, A3) microsparitic dolomite zoom, B1) pseudospherulitic dolomite, B2)
JSR

pseudospherulitic dolomite, B3) pseudospherulitic dolomite, C1) coarse cloudy dolomite with dissolved core, C2) coarse cloudy dolomite, C3) coarse cloudy dolomite.
JSR PERSPECTIVES 257

TABLE 2.— Summary of fluid-inclusion data.

Average Homogenization Homogenization Temperature


Dolomite Types Number of Samples Temperature (uC) (uC) Range Std. Dev.
Pseudospherulitic 8 84.1 77 to 91 5
Clear coarse cement 43 87.9 61 to 131 13.2
Coarse cloudy 21 88.8 81 to 111 7.2
Average Final Melting
Dolomite Types Number of Samples Temperature for ice (Tm ice) (uC) Range of (Tm ice) (uC) Std. Dev.
Pseudospherulitic 8 226.3 231 to 220 4.3
Clear coarse cement 28 225.7 233 to 221 3.9
Coarse cloudy 11 225.7 231 to 221 3.8
Dolomite Types Number of Samples Average weight % salts Range weight % salts Std. Dev.
Pseudospherulitic 8 25.1 21.7 to 26.3 1.9
Clear coarse cement 28 24.8 22.3 to 25.9 1.67
Coarse cloudy 11 24.8 22.3 to 26.3 1.33

CaCl2–H2O brine. Given the assumptions, the mean Tm ice yields a dolomite. It is in nodules and is spatially associated with anhydrite
salinity of 25.1 weight percent salt (Fig. 9; Table 2). nodules. Crystals are polymodal, range in size from 0.1 mm to 2.5 mm
(very coarsely crystalline), and consist of equant, planar, euhedral to
Clear Coarse Dolomite (Cement).—Clear coarse dolomite cement subhedral dolomite rhombs present in a hypidiotopic mosaic. A small
occurs throughout the study interval and constitutes approximately percentage (, 10%) of these crystals have cores that have been dissolved
12% of the total dolomite present. Dolomite cement fills and lines pores and remain open today. These dissolved cores may represent relatively
and cracks associated with rhizoconcretions. Two generations of cement recent dissolution of the dolomite. Cloudy and clear zones alternate.
are distinguished on the basis of grain size, shape, crystal color, clarity, Cloudy zones contain abundant fluid inclusions (4 to 20 mm) whereas
and location in the pore. The first generation typically lines the inner clear zones contain only a few fluid inclusions (Fig. 7G). These crystals
parts of pores and consists of finely crystalline (15 to 30 mm), unimodal exhibit moderate to bright, orange and red luminescence and an overall
crystals exhibiting a drusy texture, consisting of equant, planar, subhedral mottled texture, with localized areas exhibiting concentric zonation.
to anhedral dolomite, in a hypidiotopic mosaic. This generation is cloudy, Inclusion-rich zones display bright luminescence. Inclusion-poor zones
tan to gray in transmitted light, and contains only rare resolvable fluid display moderate luminescence (Fig. 7H). Microprobe analysis indicates
inclusions. Cathodoluminescence reveals moderate orange and red that the coarse cloudy dolomite is relatively calcic (45.6% Mg and 54.4%
luminescence with a mottled texture. Ca; Table 1) with low Sr (BDL) and Na (658) and elevated Fe
The second generation lines the outer rims ofpores and is medium (80 mm) to (41433 ppm) and Mn (5206) concentrations (Table 1).
coarsely crystalline (300 mm), unimodal crystals exhibiting a blocky texture The coarse cloudy dolomite phase has inclusion-rich zones located in
consisting of equant, planar, euhedral dolomite rhombs in a hypidiotopic specific growth zones (cloudy zones; Fig. 8C). These inclusions may have
mosaic (Fig. 7E). Crystals are typically clear in transmitted light with been trapped during initial dolomite growth, or during later recrystalli-
moderately abundant fluid inclusions (2 to 10 mm). In cathodoluminescence, zation. Inclusions are 2 to 10 mm, two-phase, and appear in two
this generation exhibits an overall mottled texture with moderate to bright dimensions with areas of approximately 25% gas phase and 75% liquid
luminescence and localized zones of bright luminescence along the outer edges phase. Fluid-inclusion analysis was conducted on five fluid-inclusion
of some crystals (Fig. 7F). Microprobe analysis indicates a relatively calcic assemblages and resulted in twenty-one homogenization temperatures
composition(mean 5 43%,Mgandmean 5 57%Ca)withlowSr(BDL),Na ranging from 81uC to 111uC with a mean value of 88.8uC (Table 2).
(mean 5 460 ppm), and elevated Fe (mean 5 21824 ppm), and Mn Eleven ice final melting temperatures were determined and ranged from
(mean 5 2838) concentrations (Table 1). 221uC to 231uC with a mean for 225.7uC; one eutectic was observed at
Inclusions are interpreted as having a primary origin owing to their 245uC. The low eutectic supports an interpretation of an NaCl–CaCl2–
localization in distinct growth zones. Inclusions may have been trapped H2O composition for the brine (Crawford 1981). Mean salinity is
during the initial growth of the crystal or during later recrystallization. calculated to be 24.8 wt. % salts (Table 2; Fig. 9).
They are typically two phase and 2 to 10 mm in size. They appear in two
dimensions and consist of 20% gas and 80% liquid (Fig. 8). Fluid- Isotopic Attributes
inclusion analysis was conducted on thirteen separate FIAs and resulted
in forty-three homogenization temperatures ranging from 61uC to 131uC, Carbon and Oxygen Isotopes.—Four paleosol profiles from varying
with a mean value of 87.9uC (Table 2). Five of the FIAs produced stratigraphic depths were selected for carbon and oxygen isotope
consistent Th, yielding Th data ranging from 61uC to 13uC. Twenty-eight analyses. The d13C values for the four paleosol profiles (Fig 10; Table 3)
ice final melting temperatures were determined and ranged from 221uC exhibit relatively invariant trends, with average values around 23.84%
to 233uC with a mean of 225.7uC; two eutectics were measured at 246uC PDB (Profile A; Fig. 10, Profile A; Table 3) and 25.3% PDB (Profile B
and 248uC. The low eutectics support an interpretation of an NaCl– and C; Fig. 10, Profiles B, C; Table 3). The d13C values exhibit marked
CaCl2–H2O composition for the brine (Crawford 1981). Mean salinity is up-profile enrichment in profile D (from 25.5% to +0.49% PDB; Fig. 10,
calculated to be 24.8 wt. % salts (Table 2; Fig. 9). Profile A; Table 3). For oxygen isotopes, profiles B and C exhibit
relatively invariant trends, and average values are between 20.3% and
Coarse Cloudy Dolomite.—Coarse cloudy dolomite occurs in the lower 20.6% (PDB; Fig. 10, Profile B, C: Table 3). Two profiles show more
half of the study interval and represents approximately 2% of the total variation, with d18O values in profile A exhibiting a slight up-profile
258 K.A. HARTIG ET AL. JSR

FIG. 9.—Fluid-inclusion data by fluid-inclusion assemblage (FIA). Samples for fluid-inclusion analysis were taken from the Libby and Pittard cores. Samples from the
Libby core consist of the coarse cloudy dolomite and the dolomite cement. Samples taken from the Pittard core consist of the pseudospherulitic dolomite. All dolomite
types exhibit similar homogenization temperatures, mostly ranging from 70uC to 110uC, with only a few outliers. The final melting temperatures for all three dolomite
types are similar and appear to cluster between temperatures of 220uC to 234uC.
JSR PERSPECTIVES 259

enrichment (d18O 5 0.0% to 2.0%), and d18O values in profile D Cañaveras 1999; Mack et al. 2000). Chafetz and Butler (1980) suggested
exhibiting marked up-profile depletion (d18O 5 +3.0% to 23.5% PDB; that pseudospherulitic textures form by the ‘‘neomorphism of Micro-
Fig. 10, Profile A, D; Table 3). codium, or Microcodium-like globular bodies,’’ whereas Rossi and
Cañaveras (1999) linked pseudospherulitic textures to solution–precipi-
Sulfur Isotopes.—In addition to the carbon and oxygen isotope data, tation processes occurring at or near the water table, under predomi-
two samples were selected for sulfur isotope measurements. The first nantly oxidizing geochemical conditions, and possibly bacterially
sample from the Nikkel #1 core at the base of the Cimarron Anhydrite mediated. Similar textures have been observed to form as siderite in
(Fig. 5) yielded two d34S values of 10.3% and 10.4%. The second sample, wetland soils (e.g., Stoops 1983; Ludvigson et al. 1998). We infer an initial
from an anhydrite nodule present near the base of the Abo–Tubb sections groundwater origin for the pseudospherulitic dolomite, but the rare
in the Libby #1 core, included two d34S values of 9.6% and 9.7%. presence of solid calcite inclusions again suggests a possible early calcite
composition, and later dolomitization.
DISCUSSION Taken together, these data indicate that groundwater processes that
occurred penecontemporaneous with deposition of the Abo–Tubb
As evinced by the macroscopic and microscopic attributes documented interval in the early Permian resulted in precipitation of pseudospherulitic
both here and in Kessler et al. (2001), paleosols are clearly present in the calcite, which was later dolomitized, as further detailed below (Fig. 11).
Abo–Tubb interval in northeastern New Mexico and all contain dolomite The concentration of this dolomite phase in the lower half of the study
as the sole carbonate phase, with the exception of calcite inclusions in the interval likely reflects groundwater fluid flow that was focused along the
dolomite. In addition to the pedogenic textures, some of the carbonate Permian–Precambrian nonconformity surface.
present in the study interval exhibits attributes typical of groundwater
precipitation and additionally bears evidence for recrystallization (see
Dolomitization (Seepage Reflux; Leonardian)
below). Although the carbonate in the study interval is now dolomite, it
may have precipitated either initially as dolomite or initially as calcite that Given the kinetic and thermodynamic barriers to precipitating primary
has subsequently been replaced by dolomite. pedogenic dolomite, and the mottled luminescence pattern and rare
presence of solid calcite inclusions in the microsparitic and pseudo-
Early (Pedogenic) Carbonate Formation spherulitic dolomite phases, we favor the interpretation that these phases
originated as calcite and were later dolomitized. Additional evidence from
Multiple macroscopic and petrographic features suggest that much of the coarser phases, together with the stratigraphic setting, offer a
the microsparitic dolomite currently present in the Abo–Tubb interval mechanism to achieve such dolomitization.
formed initially as pedogenic carbonate (Fig. 11). Primary evidence for a
The Leonardian Cimarron Anhydrite directly overlying the Abo–Tubb
pedogenic origin of the carbonate in the Abo–Tubb interval is the
interval is likely a shallow evaporitic marine or mixed marine–nonmarine
presence of typical pedogenic textures. For example, much of the micrite
evaporite unit (Hylton 1988; Hovorka et al. 1993). The conditions that led
occurs in the form of displacive nodules and vertically aligned tubules
to the deposition of the Cimmaron Anhydrite could have induced reflux
(rhizoliths) concentrated in discrete (inferred Bk) horizons. Furthermore,
of Mg-rich (marine or mixed marine–nonmarine) brines through the
these horizons also exhibit ped structures (locally with slickensides), and
Abo–Tubb study interval. Oxygen and sulfur isotope data are consistent
microsparitic dolomite occurs in the finest-grained (muddy) intervals.
with reflux of marine-influenced brines. Sulfur isotopes from both the
These observations are in contrast to groundwater-associated carbonate,
Cimarron Anhydrite and an anhydrite nodule near the base of the study
which typically occurs in coarser-grained, more permeable intervals
interval contain similar d34S values ranging from 9.6% to 10.4%,
(Pimental et al. 1996; Mack et al. 2000). Homogeneous microsparitic
dolomite predominates and commonly exhibits circumgranular cracking suggesting that they were precipitated from a fluid of similar origin. These
and displacive textures consistent with pedogenic carbonate formation d34S values are close to early Permian seawater values (, 12%; Claypool
(Pimental et al. 1996; Mack et al. 2000). Calcite inclusions suggest the et al. 1980; Strauss 1997). Oxygen isotopes also provide evidence for a
possibility that the dolomicrite formed initially as pedogenic calcite and possible low-temperature, evaporated-fluid event of dolomitization.
was later dolomitized. However, it is also possible that these inclusions Oxygen isotopes exchange readily during replacement, and so the current
reflect dedolomitization of original pedogenic dolomite. Although it d18O values exhibited in the paleosol profiles were likely altered during
remains possible that the dolomite formed as a primary pedogenic phase, initial dolomitization and later dolomite recrystallization (cf. Mora and
because other phases record evidence for later diagenetic dolomitization Driese 1999). The most positive d18O values present in the dolomite of the
events (discussed below), a simpler explanation is that these diagenetic Abo–Tubb interval likely approach values closer to initial dolomite
events similarly induced wholesale dolomitization of early pedogenic formation. These values cluster around 3%, and assuming about 3%
calcite. enrichment for dolomite relative to calcite (Land 1980; McKenzie et al.
Carbon isotope values from the dolomite nodules and rhizoconcretions 1995) the values could reflect original values substantially enriched
in profile D could reflect a pedogenic trend, wherein d13C becomes relative to Permian seawater calcite values of 23.47 and 22.63% (Rao
enriched up-profile, owing to the interaction of relatively enriched and Green 1982; Rao 1988; Grossman 1994). With 3% dolomite
atmospheric carbon with relatively depleted organic carbon (Cerling et al. fractionation added, seawater dolomite would be expected around 20.5
1989). However, the positive correlation between d13C and d18O in two of to +0.5% The most positive oxygen values for dolomite are several per
the profiles suggests physical mixing of multiple diagenetic phases. The mil more positive than those predicted from normal Permian seawater,
overall negative d13C compositions could reflect preservation of the and are therefore consistent with dolomite precipitation from evaporated
original pedogenic signature despite probable later alteration. The d13C Permian seawater or mixed marine–nonmarine water. These data are
values are consistent with values reported for paleosol carbonate consistent with the reflux interpretation. The more negative values likely
(Goldstein 1991; Mora and Driese 1999; Wright and Vanstone 1991). reflect diagenetic resetting at higher temperature.

Early Groundwater Carbonate Formation Late Recrystallization (Cenozoic)


Some authors have associated psuedospherulitic textures with low- The coarse and cloudy dolomite cement and the pseudospherulitic
temperature groundwater processes (e.g., Pimental et al. 1996; Rossi and dolomite have mottled fabrics in CL, indicating recrystallization of an
260 K.A. HARTIG ET AL. JSR
JSR PERSPECTIVES 261

TABLE 3.— Summary of isotopic data.

Paleosol Number of
Profile Samples mean d13C d13C Std. Dev. mean d18O range d18O Std. Dev.
A 15 23.84 22.42 to 24.35 0.45 1.52 20.58 to 2.54 0.9
B 13 25.24 25.38 to 25.06 0.1 20.34 20.5 to .008 0.15
C 14 25.32 25.47 to 25.19 0.1 20.60 20.9 to 20.23 0.22
D 11 24.29 25.37 to 0.49 1.79 1.46 22.56 to 2.85 1.59

earlier dolomite and a later dolomite-precipitating event (Fig. 11). Areas Aasm 1997; White and Al-Aasm 1997; Reinhold 1998; Lonnee and Al-
with the brightest luminescence include inclusion-rich zones, mottles, Aasm 2000; Keller et al. 2000; Cantrell et al. 2004; Smith and Davies
fracture fillings, and the outermost concentric growth zone, suggesting a 2006; Davies and Smith 2006).
change in fluid conditions after initial crystal growth and fracturing that Multiple tectonic and magmatic events have affected the region within
led to cementation and recrystallization. and around northeastern New Mexico since the deposition of the Abo–
Cloudy and clear, coarse dolomite cement has elevated levels of Fe Tubb interval. Specifically, regional tectonics associated with both the
and Mn relative to the microsparitic and pseudospherulitic dolomite. Mesozoic Sevier Orogeny and Laramide Orogeny affected the greater
The elevated Fe and Mn levels are consistent with hydrothermal fluids southern Rocky Mountain region, and more specifically the Raton Basin
(Schrijver et al. 1996) and deep basinal brines (Yoo et al. 2000), and region (Baltz 1965; Martinez 1989; Woodward 1987; Woodward et al.
such elevated levels clearly record precipitation under reducing 1999). Dolomite-precipitating basinal fluid flow has been documented
conditions. extensively in relation to regional tectonic events; fluids derived from
FIAs with consistent Th values (sensu Goldstein and Reynolds 1994) adjacent deep basins coupled with faulting, uplift, permeable stratigraphic
yield homogenization temperatures between 70uC and 110uC, evidence units, and a basal unconformity could provide for the necessary pathways
that high temperatures existed during dolomite cementation and for fluid migration (Gao et al. 1992a; Gao et al. 1992b; Gao et al. 1995;
dolomite recrystallization. FIAs with inconsistent Th yielded values as Oliver 1992, 1996). In addition, late Cenozoic magmatism and extrusion
high as 131uC, indicating that the Abo–Tubb interval reached that have been associated with the Spanish Peaks area (southern Colorado),
temperature after the dolomite existed. The fluid-inclusion eutectic the Raton–Clayton volcanics, and the Rio Grande Rift (e.g., Manely
temperatures between 243uC and 248uC indicate abundant divalent 1984; Calvin 1987; Woodward 1987; Bove et al. 1995). Possible migration
cations. Final melting temperatures for ice between 233uC and 220uC, of hydrothermal fluids into the study area in association with these events
indicate approximately 24 weight percent salt. Thus, high-temperature is reasonable. The hydrothermal conditions recorded by fluid inclusions
fluids that caused dolomite cementation and dolomite recrystallization in the dolomite, however, likely predated charging of CO2 into the Bravo
were highly saline and had chemistries other than simple NaCl. The Dome reservoir. No clathrates were documented in the fluid inclusions,
FIAs with consistent Th (70uC and 110uC) and those from inconsistent and no CO2 fluid inclusions were observed, suggesting that CO2 had not
Th (up to 131uC), are higher than the relatively low burial temperatures yet charged the reservoir during the late events of dolomite precipitation
estimated previously for the Abo–Tubb interval (, 50uC; Kessler et al. and recrystallization. This observation is consistent with previous
2001). Upper Permian through Mesozoic overburden was likely , 300 interpretations that the CO2 charging was associated with recent basaltic
to 1000 m (Broadhead 1990). Assuming a geothermal gradient of 25uC/ magmatism in the region, between 1 Ma and 0.8 ka (Broadhead 1998;
km, the maximum burial temperature of approximately 50uC is Baines and Worden 2004; Broadhead and Gillard 2004; Cassidy 2005).
inconsistent with fluid-inclusion Th values as high as 131uC. Thus,
the system was either conductively heated during a time of higher heat Implications for CO2 Sequestration
flow or was subject to injection of warmer fluids into cooler rocks.
Systems with injection of aqueous fluids that are appreciably warmer Geologic sequestration of CO2 in subsurface saline aquifers has become
than ambient (5uC or greater; White 1957) are defined as hydrothermal, an active area of research as governments consider the effects of increased
and are worthy of consideration for recrystallizing and precipitating atmospheric CO2 on climate change. Baines and Worden (2004) have
dolomite. studied, as a natural analog for CO2 sequestration, several aquifer
systems in which the natural impact of CO2 addition could be evaluated.
Higher-Temperature Fluids The Bravo Dome is one such system where simple geochemical models
predicted that unstable silicates (e.g., calcic plagioclase, calcic zeolites,
Dolomite formation induced by migration of hydrothermal fluids is clays) would dissolve and lead to carbonate precipitation after addition of
well documented (e.g., Gregg 1985; Schrijver et al. 1996; Durocher and CO2 and its dissolution into the saline aquifer. Their findings were, on the
Al-Aasm 1997; White and Al-Aasm 1997; Reinhold 1998; Lonnee and Al- contrary, that the aluminosilicate reactions were likely too slow to
Aasm 2000; Keller et al. 2000; Machel and Lonnee 2002). Dolomite produce the predicted results, and that carbonate dissolution would be
recrystallization and precipitation by hydrothermal fluids in the Abo– the expected reaction. Our study confirms that late-stage dissolution of
Tubb interval requires a mechanism for upward fluid movement, which minor dolomite phases are in fact the dominant mineralogical result of
could be related to deformation, volcanism in surrounding regions, or CO2 injection, and that it is likely that the aluminosilicate reactions are
both (Leach and Sangster 1993; Wojcik et al. 1997; Durocher and Al- too slow to buffer the system.

FIG. 10.—Paleosol profiles and associated isotopic values for paleosol profiles. All isotopic data are reported in per mil relative to VPDB. Profiles A, B, and C exhibit
relatively consistent d13C and d18O values throughout the profile. Conversely, Profile D exhibits significant changes in d13C and d18O values near the top of the
paleosol profile.
262 K.A. HARTIG ET AL. JSR

FIG. 11.—Summary of paragenetic sequence


in the Abo–Tubb interval, outlining the various
events that have contributed to the carbonate
mineral paragenesis.

CONCLUSIONS tectonism (Fig. 12). This example of high-temperature alteration of low-


temperature dolomite is useful in evaluating controversies about the
Lower Permian Abo–Tubb subsurface strata of northeastern New
origin of dolomite in hydrocarbon reservoirs with hydrothermal
Mexico consist of 30 to 120 m of fluvial and eolian deposits containing
overprints.
numerous dolomitic paleosols. Macroscopic characteristics of much of
Because the Bravo Dome system was initially charged with CO2 during
the dolomite appear pedogenic, but petrographic observations and fluid-
inclusion data coupled with microprobe data and isotopic data provide the last 1 Ma to 8 ka, the paragenesis is useful in testing hypotheses of
evidence for dolomitization and recrystallization of the coarser phases in long-term mineral transformations associated with CO2 sequestration.
particular. Although we cannot completely eliminate the possibility of a Given an aquifer of arkosic sandstone and minor carbonate phases, the
primary pedogenic origin for the volumetrically predominant micro- results show that the dominant initial mineral reaction after CO2 injection
sparitic dolomite phase, the most reasonable explanation is that the will be dissolution of carbonate minerals. Aluminosilicate reactions
dolomite currently present in the Abo–Tubb interval was formed by a appear to be too slow to buffer the system and do not initially lead to
series of later events. Implications of this are significant because dolomite carbonate precipitation.
in paleosols has been interpreted, at times, as an indicator of climate or
atmospheric conditions during pedogenesis. If dolomite forms by later ACKNOWLEDGMENTS
replacement of paleosol calcite, then its presence is irrelevant for
This work represents research conducted for Hartig’s M.S. thesis under the
interpreting paleoenvironment during soil formation. Hence, these results
guidance of Dr. Gerilyn Soreghan, Dr. Michael Engel, Dr. Thomas Dewers,
reiterate the importance of careful diagenetic study of minerals in and Dr. Robert Goldstein. This research was partially supported by the
paleosols, to disentangle primary from replaced phases. donors of the Grants in Aid Foundation, administered by the American
After pedogenic and groundwater calcite formed during Early Permian Association of Petroleum Geologists. Acknowledgment is made to the donors
deposition (Wolfcampian–Leonardian), calcite was dolomitized in of the Petroleum Research Fund 35217-AC8, administered by the American
association with low-temperature seepage reflux, during Early Permian Chemical Society, for partial support of this research. Special thanks to Dr.
deposition of the overlying Cimarron Anhydrite (Leonardian). This Early Mike Soreghan for technical assistance and guidance throughout this project
and to Dr. Kathleen Counter Benison for her guidance and mentorship. We
Permian dolomite was extensively recrystallized, and dolomite cement would also like to thank Eugene Rankey, Ronald Broadhead, and David
precipitated from highly saline waters at temperatures well above those Barnes for their insightful comments and reviews regarding this paper.
that can be ascribed to normal burial heating. This alteration was mostly Analytical and technical support at the University of Oklahoma was provided
likely caused by hydrothermal fluid flow associated with Cenozoic by R. Maynard and M. Engel (stable isotopes) and G. Morgan (microprobe).
JSR PERSPECTIVES 263

FIG. 12.—Conceptual model of dolomite for-


mation in the Abo–Tubb interval. A) Diagram
outlining the setting where pedogenic calcite
forms during deposition. B) Diagram outlining
the process of saline fluids seeping into the
underlying sediment and dolomitizing the pre-
cursor pedogenic calcite during deposition of the
Cimarron Anhydrite. C) Diagram outlining
recrystallization of dolomite during Cenozoic
influx of hydrothermal fluids.

REFERENCES BALTZ, E.H., 1965, Stratigraphy and history of Raton Basin and notes on San Luis
Basin, Colorado–New Mexico: American Association of Petroleum Geologists,
ALONSO-ZARZA, A.M., 2003, Palaeoenvironmental significance of palustrine carbonates Bulletin, v. 49, p. 2041–2075.
and calcretes in the geological record: Earth-Science Reviews, v. 60, p. 261–298. BARNETT, A.J., WRIGHT, V.P., AND CROWLEY, S.F., in press, Recognition and
ARMENTEROS, I., ANGELES BUSTILLIO, M.A., AND BLANCO, J.A., 1995, Pedogenic and significance of paludal dolomites: Late Mississippian, Kentucky, USA, in Morad,
groundwater processes in a closed Miocene basin (northern Spain): Sedimentary S., and Ketzer, M., eds., Linking Diagenesis to Sequence Stratigraphy of Sedi-
Geology, v. 99, p. 17–36. mentary Rocks: International Association of Sedimentologists, Special Publi-
BACHMAN, G.D., AND MACHETTE, M.N., 1977, Calcic soils and calcretes in the cation.
southwestern United States: U.S. Geological Survey, Open-File Report 77-794, 163 p. BENITO, M.I., DE LA HORRA, R., BARRENECHEA, J.F., LÓPEZ-GÓMEZ, J., RODAS, M.,
BAINES, S.J., AND WORDEN, R.H., 2004, The long-term fate of CO2 in the subsurface: ALONSON-AZCARATE, J., ARCHE, A., AND LUQUE, J., 2005, Late Permian continental
natural analogues for CO2 storage, in Baines, S.J., and Worden, R.H., eds., sediments in the SE Iberian Ranges, eastern Spain: Petrological and mineralogical
Geological Storage of Carbon Dioxide: Geological Society of London, Special characteristics and palaeoenvironmental significance: Palaeogeography, Palaeoclima-
Publication 233, p. 59–85. tology, Palaeoecology, v. 2291–2, p. 24–39.
BAKKER, R.J., 2003, Package FLUIDS 1. Computer programs for analysis of fluid BOTHA, G.A., AND HUGHES, J.C., 1992, Pedogenic palygorskite and dolomite in a late
inclusion data and for modelling bulk fluid properties: Chemical Geology, v. 194, Neogene sedimentary succession, northwestern Transvaal, South Africa: Geoderma,
p. 3–23. v. 53, p. 139–154.
264 K.A. HARTIG ET AL. JSR
BOVE, D.J., RATTÉ, J.C., MCINTOSH, W.C., SNEE, L.W., AND FUTA, K., 1995, The GOLDSTEIN, R.H., 2003, Petrographic analysis of fluid inclusions, in Samson, I.,
evolution of Eagle Peak volcano–a distinct phase of Middle Miocene volcanism in the Anderson, A., and Marshall, D., eds., Fluid Inclusions: Analysis and Interpretation:
Western Mogollon–Datil volcanic field, New Mexico: Journal of Volcanology and Mineralogical Association of Canada, Short Course 32, p. 9–54.
Geothermal Research, v. 69, p. 159–186. GOLDSTEIN, R.H., AND REYNOLDS, T.J., 1994. Systematics of Fluid Inclusion in
BROADHEAD, R.F., 1990, Bravo dome carbon dioxide field, in Beaumont, E.A., and Diagenetic Minerals: SEPM, Short Course 31, 199 p.
Foster, N.H., eds., Structural Traps I, Tectonic Fold Traps: American Association of GREGG, J.M., 1985, Regional epigenic dolomitization in the Bonneterre dolomite
Petroleum Geologists, Treatise of Petroleum Geology, Atlas of Oil and Gas Fields, (Cambrian), southern Missouri: Geology, v. 13, p. 503–506.
p. 213–232. GREGG, J.M., AND SHELTON, K.L., 1990, Dolomitization and dolomite neomorphism in
BROADHEAD, R.F., 1993, Carbon dioxide in northeast New Mexico: West Texas the back-reef facies of the Bonneterre and Davis formations (Cambrian), southeastern
Geological Society, Bulletin, v. 32, p. 5–8. Missouri: Journal of Sedimentary Petrology, v. 60, p. 549–562.
BROADHEAD, R.F., 1998, Natural accumulations of carbon dioxide in the New Mexico GROSSMAN, E.L., 1994, The carbon and oxygen isotope record during the evolution of
region—Where are they, how do they occur, and what are the uses for CO2?: Lite Pangea: Carboniferous to Triassic, in Klein, G.D., ed., Pangea: Paleoclimate,
Geology, no. 20, p. 2–7. Tectonics, and Sedimentation during Accretion, Zenith, and Breakup
BROADHEAD, R.F., AND GILLARD, L., 2004, Helium in New Mexico: Geologic of a Supercontinent: Geological Society of America, Special Paper 288, p. 207–228.
Distribution and Exploration Possibilities: New Mexico Bureau of Geology and HARDIE, L.A., 1987, Dolomitization: a critical view of some current views: Journal of
Mineral Resources, Open File Report, 483, 62 p. Sedimentary Petrology, v. 57, p. 166–183.
BROADHEAD, R.F., AND KING, W.E., 1988, Petroleum Geology of Pennsylvanian and HOVORKA, S.D., KNAUTH, L.P., FISHER, R.S., AND GAO, G., 1993, Marine to nonmarine
Lower Permian Strata, Tucumcari Basin, East-Central New Mexico: New Mexico facies transition in Permian evaporites of the Palo Duro Basin, Texas: geochemical
Bureau of Mines and Mineral Resources, Bulletin, 119, 75 p. response: Geological Society of America, Bulletin, v. 104, p. 1119–1134.
CALVIN, E.M., 1987, A review of the volcanic history and stratigraphy of northeastern HYLTON, K.A., 1988, Environment of deposition of the Stone Corral Formation (Lower
New Mexico, the Ocate and Raton–Clayton Volcanic field, in Lucas, S.G., and Hunt, Permian) Kansas: The Compass, v. 65, p. 70–79.
A.P., eds., Northeastern New Mexico: New Mexico Geological Society, Guidebook KELLER, T.J., GREGG, J.M., AND SHELTON, K.L., 2000, Fluid migration and associated
38, p. 83–86. diagenesis in the Greater Reelfoot rift region, Midcontinent, United States:
CANTRELL, D.L., SWART, P.K., AND HAGERTY, R.M., 2004, Genesis and characterization Geological Society of America, Bulletin, v. 112, p. 1680–1693.
of dolomite, Arab-D reservoir, Ghawar field, Saudi Arabia: GeoArabia, v. 9, p. 1–26. KENWARD, P.A., GOLDSTEIN, R.H., GONZALEZ, L.A., AND ROBERTS, J.A., 2009,
CAPO, R.C., WHIPKEY, C.E., BLACHERE, J., AND CHADWICK, O.A., 2000, Pedogenic origin Precipitation of low-temperature dolomite from an anaerobic microbial consortium:
of dolomite in a basaltic weathering profile, Kohala Peninsula, Hawaii: Geology, v. 83, the role of methanogenic Archaea: Geobiology, v. 7, p. 1–10.
p. 515–524. KESSLER, J., SOREGHAN, G., AND WACKER, J., 2001, Equatorial aridity in western Pangea:
CARPENTER, A.B., TROUT, M.L., AND PICKETT, E.E., 1974, Preliminary report on the Lower Permian loessite and dolomitic Paleosols in northeastern New Mexico, U.S.A.:
origin and chemical evolution of lead and zinc-rich oil field brines in central Journal of Sedimentary Research, v. 71, p. 817–832.
Mississippi: Economic Geology, v. 69, p. 1191–1206. KOHUT, C.K., MUEHLENBACHS, K., AND DUDAS, M.J., 1995, Authigenic dolomite in
CASSIDY, M.M., 2005. Occurrence and origin of free carbon dioxide gas deposits in the saline soils in Alberta, Canada: American Journal of Soil Science, v. 59, p. 1499–1504.
earth’s continental crust [Ph.D. dissertation]: University of Houston, 242 p. LAND, L.S., 1980, The isotopic and trace element geochemistry of dolomite: the state of
CERLING, T.E., QUADE, J., WANG, Y., AND BOWMAN, J.R., 1989, Carbon isotopes in soils the art, in Zenger, D.H., Dunham, J.B., and Ethington, R.L., eds., Concepts and
and paleosols as ecological and paleoecologic indicators, Nature, v. 341, p. 138–139. Models of Dolomitization: SEPM, Special Publication 28, p. 87–110.
CHAFETZ, H.S., AND BUTLER, J.C., 1980, Petrology of recent caliche, pisolites, LEACH, D.L., AND SANGSTER, D.F., 1993, Mississippi Valley–type lead–zinc deposits, in
spherulites, and speleothem deposits from central Texas: Sedimentology, v. 27, Kirkham, R.V., Sinclair, W.D., Thorp, R.I., and Duke, J.M., eds., Mineral Deposit
p. 497–518. Modeling: Geological Association of Canada, Special Paper 40, p. 289–314.
CLAYPOOL, G.E., HOLSER, W.T., KAPLAN, I.R., SAKAI, H., AND ZAK, I., 1980, The age LONNEE, J., AND AL-AASM, I.S., 2000, Dolomitization and fluid evolution in the Middle
curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation: Devonian Sulphur Point Formation, Rainbow South field, Alberta: Petrographic and
Chemical Geology, v. 28, p. 199–260. geochemical evidence: Bulletin of Canadian Petroleum Geology, v. 48, p. 262–283.
COLOSON, J., AND COJAN, I., 1996, Groundwater dolocretes in a lake-marginal LUDVIGSON, G.A., GONZALEZ, L.A., METZGER, R.A., WITZKE, B.J., BRENNER, R.L.,
environment; an alternative model for dolocrete formation in continental setting MURILLO, A.P., AND WHITE, T.S., 1998, Meteoric sphaerosiderite lines and their use
(Danian of Providence Basin, France: Sedimentology, v. 43, p. 175–188. for paleohydrology and paleoclimatology: Geology, v. 26, p. 1039–1042.
CRAWFORD, M.L., 1981, Phase equilibria in aqueous fluid inclusions, in Hollister, L.S., MACHEL, H.G., 2004. Concepts and models of dolomitization: a critical reappraisal:
and Crawford, M.L., eds., Fluid Inclusions: Applications to Petrology: Mineralogical Geological Society of London, Special Publication 235, p. 7–63.
Association of Canada, Short Course Handbook, v. 6, p. 75–100. MACHEL, H.G., AND LONNEE, J., 2002, Hydrothermal dolomite—a product of poor
DAVIES, G.R., AND SMITH, T., 2006, Structurally controlled hydrothermal dolomite definition and imagination: Sedimentary Geology, v. 152, p. 163–171.
reservoir facies: an overview: American Association of Petroleum Geologists, Bulletin, MACHETTE, M.N., 1985, Calcic soils of southwestern United States, in Weide, D.L., ed.,
v. 90, p. 1641–1690. Soils and Quaternary Geology of Southeastern United States: Geological Society of
DUROCHER, S., AND AL-AASM, I.S., 1997, Dolomitization and neomorphism of America, Special Paper 203, p. 1–21.
Mississippian (Visean) upper Debolt Formation, Blueberry Field, northeastern MACK, G.H., JAMES, W.C., AND MONGER, H.C., 1993, Classification of paleosols:
British Columbia: geologic, petrologic, and chemical evidence: American Association Geological Society of America, Bulletin, v. 105, p. 129–136.
of Petroleum Geologists, Bulletin, v. 81, p. 954–977. MACK, G.H., COLE, D.R., AND TREVIÑO, L., 2000, The distribution and discrimination of
EL-SAYED, M.I., FAIRCHILD, I.J., AND SPIRO, B., 1991, Kuwaiti dolocretes: petrology, shallow, authigenic carbonate in the Pliocene–Pleistocene Palomas Basin, southern
geochemistry, and groundwater origin: Sedimentary Geology, v. 73, p. 59–75. Rio Grande rift: Geological Society of America, Bulletin, v. 112, p. 643–656.
GARCES, B.L.V., AND AGUILAR, J.G., 1994, Permian saline lakes in the Aragon–Bearn MANELY, K., 1984, Brief summary of the Tertiary geologic history of the Rio Grande
basin, Western Pyrenees, in Renaut, R.W., and Last, W.M., eds., Sedimentology and Rift in Northern New Mexico, in Baldridge, W.S., Dickerson, P.W., Riecker, R.E.,
Geochemistry of Modern and Ancient Saline Lakes: SEPM, Special Publication 50, and Zidek, J., eds., Rio Grande Rift: Northern New Mexico: New Mexico Geological
p. 267–290. Society, Guidebook 35, p. 63–66.
GAO, G., AND LAND, L.S., 1991, Early Ordovician Cool Creek Dolomite, middle MARTINEZ, R., 1989, Summary of Laramide Orogeny in New Mexico, in , LORENZ, J.C.,
Arbuckle Group, Slickhills, SW Oklahoma, USA: origin and modification: Journal of AND LUCAS, S.G., 1989, Energy Frontiers in the Rockies: Albuquerque Geological
Sedimentary Petrology, v. 61, p. 161–173. Society, Albuquerque, New Mexico, p. 171–176.
GAO, G., ELMORE, R.D., AND LAND, L.S., 1992a, Geochemical constraints on the origin MCKENZIE, J.A., HSU, K.J., AND SCHNEIDER, J.F., 1980, Movement of subsurface water
of calcite veins and associated limestone alteration, Ordovician Viola Group, under the sabkha, Abu Dhabi, in Zenger, D.H., Dunham, J.B., and Ethington, R.L.,
Arbuckle Mountains, Oklahoma, U.S.A.: Chemical Geology, v. 98, p. 257–269. eds., Concepts and Models of Dolomitization: SEPM, Special Publication 28,
GAO, G., LAND, L.S., AND FOLK, R.L., 1992b, Meteoric modification of early dolomite p. 11–30.
and late dolomitization by basinal fluids, Upper Arbuckle Group, Slick Hills, MCKENZIE, J.A., TERANES, J., AND FISCHER, A., 1995, Oxygen isotope records of abrupt
Southwestern Oklahoma: American Association of Petroleum Geologists, Bulletin, (decades) continental climate change in late glacial/early Holocene and modern Swiss
v. 76, p. 1649–1664. lacustrine carbonate sequences: EOS, Transactions, American Geophysical Union,
GAO, G., LAND, L.S., AND ELMORE, R.D., 1995, Multiple episodes of dolomitization in 1995 Spring Meeting, v. 76, 282 p.
the Arbuckle Group, Arbuckle Mountains, south-central Oklahoma: field, petro- MONTAÑEZ, I.P., AND READ, J.K., 1992, Fluid–rock interaction history during
graphic, and geochemical evidence: Journal of Sedimentary Research, v. 65, stabilization of early dolomites, upper Knox Group (lower Ordovician), U.S.
p. 321–331. Appalachians: Journal of Sedimentary Petrology, v. 62, p. 753–778.
GILE, L.H., PETERSON, F.F., AND GROSSMAN, R.B., 1966, Morphological and genetic MORA, C.I., AND DRIESE, S.G., 1999, Palaeoenvironment, paleoclimate and stable
sequences of carbonate accumulation in desert soils: Soil Science, v. 101, p. 347–360. carbon isotopes of Paleozoic red-bed paleosols, Appalachian Basin, USA and
GILFILLAN, S.M.V., BALLENTINE, C.J., HOLLAND, G., BLAGBURN, D., LOLLAR, B.S., Canada, in Thiry, M., and Simon-Coinçon, R., eds., Palaeoweathering, Palaeosur-
STEVENS, S., SCHOELL, M., AND CASSIDY, M., 2008, The noble gas chemistry of natural faces and Related Continental Deposits: International Association Sedimentologists,
CO2 gas reservoirs from the Colorado Plateau and Rocky Mountain provinces, USA: Special Publication 27, p. 61–84.
Geochimica et Cosmochimica Acta, v. 72, p. 1174–1198. MOREIRA, N.F., WALTER, L.M., VASCONCELOS, C., MCKENZIE, J.A., AND MCCALL, P.J.,
GOLDSTEIN, R.H., 1991, Stable isotope signatures associated with paleosols, Pennsyl- 2004, Role of sulfide oxidation in dolomitization; sediment and pore-water
vanian Holder Formation, New Mexico: Sedimentology, v. 38, p. 67–77. geochemistry of a modern hypersaline lagoon system: Geology, v. 32, p. 701–704.
JSR PERSPECTIVES 265

OAKES, C.K., BODNAR, R.J., AND SIMONSON, J.M., 1990, The system NaCl2–H2O the ice STOOPS, G., 1983, SEM and light microscopic observations of minerals in bog-ores of
liquids at 1atm total pressure: Geochemica et Cosmochimica Acta, v. 54, p. 603–610. Belgian Campine: Geoderma, v. 30, p. 179–186.
OLIVER, J., 1992, The spots and stains of plate tectonics: Earth-Science Reviews, v. 32, STAUDACHER, T., 1987, Upper mantle origin for Harding County well gases: Nature,
p. 77–106. v. 325, p. 605–607.
OLIVER, N.H.S., 1996, Review and classification of structural controls on fluid flow STRAUSS, H., 1997, The isotopic composition of sedimentary sulfur through time:
during regional metamorphism: Journal of Metamorphic Geology, v. 14, p. 477–492. Palaeogeography, Palaeoclimatology, Palaeoecology, v. 132, p. 97–118.
PEARCE, J.M., HOLLOWAY, S., WACKER, H., NELIS, M.K., ROCHELLE, C., AND BATENAN, VAN LITH, Y., WATHMANN, R., VASCONCELOS, C., AND MCKENZIE, J.A., 2003, Sulphate-
K., 1996, Natural occurrences as analogues for the geological disposal of carbon reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite forma-
dioxide: Energy Conversion and Management, v. 37, p. 1123–1128. tion: Geobiology, v. 1, p. 71–79.
PIMENTEL, N.L., 2002, Pedogenic and early diagenetic processes in Paleogene alluvial fan VASCONCELOS, C., MCKENZIE, J.A., BERASCONI, S., GRULIC, D., AND TEIN, A.J., 1995,
and lacustrine deposits from the Sado Basin (S Portugal): Sedimentary Geology, Microbial mediation as a possible mechanism for natural dolomite formation at low
v. 148, p. 123–138. temperatures: Nature, v. 377, p. 202–221.
PIMENTEL, N.L., WRIGHT, V.P., AND AZEVEDO, T.M., 1996, Distinguishing early WACEY, D., WRIGHT, D.T., AND BOYCE, A.J., 2007, A stable isotope study of microbial
groundwater alteration affects from pedogenesis in ancient alluvial basins: examples dolomite formation in the Coorong Region, South Australia: Chemical Geology,
from the Paleogene of southern Portugal: Sedimentary Geology, v. 105, p. 1–10. v. 244, p. 155–174.
POUCHOU, J.L., AND PICHOIR, F., 1985, ‘‘PAP’’ w(r-z) correction procedure for improved WANAS, H.A., 2002, Petrography, geochemistry and primary origin of spheroidal
quantitative microanalysis, in Armstrong, J.T., ed., Microbeam Analysis: San dolomite from the Upper Cretaceous/Lower Tertiary Maghra El-Bahari Formation
Francisco, San Francisco Press, p. 104–106. at Gabal Ataqu, northwest Gulf of Suez, Egypt: Sedimentary Geology, v. 151,
RAO, R.C., AND GREEN, D.C., 1982, Oxygen and carbon isotopes of Early Permian cold- p. 211–224.
water carbonates, Tasmania, Australia: Journal of Sedimentary Petrology, v. 52, WHIPKEY, C.E., AND HAYOB, J.L., 2008, Textural and compositional evidence for the
p. 1111–1125. evolution of pedogenic calcite and dolomite in a weathering profile on the Kohala
RAO, R.C., 1988, Oxygen and carbon isotope composition of cold-water Berriedale Peninsula, Hawaii: Carbonates and Evaporites, v. 23, p. 104–112.
Limestone (Lower Permian), Tasmania, Australia: Sedimentary Geology, v. 60, WHIPKEY, C.E., CAPO, R.C., HSIEH, J.C.C., AND CHADWICK, O.A., 2002, Development of
p. 221–231. magnesian carbonates in Quaternary soils on the Island of Hawaii: Journal of
REINHOLD, C., 1998, Multiple episodes of dolomitization and dolomite recrystallization Sedimentary Research, v. 72, p. 158–165.
during shallow burial in the Upper Jurassic shelf carbonates: eastern Swabian Alb, WHITE, D.E., 1957, Thermal waters of volcanic origin: Geological Society of America,
southern Germany: Sedimentary Geology, v. 121, p. 71–95. Bulletin, v. 68, p. 1637–1658.
ROBERTS, J.A., BENNETT, P.C., GONZÁLEZ, L.A., MACPHERSON, G.L., AND MILLIKEN, WHITE, T., AND AL-AASM, I.S., 1997, Hydrothermal dolomitization of the Mississippian
K.L., 2004, Microbial precipitation of dolomite in methanogenic groundwater: Upper Debolt Formation, Sikanni gas field, northern British Columbia, Canada:
Geology, v. 32, p. 277–280. Bulletin of Canadian Petroleum Geology, v. 45, p. 297–316.
WOJCIK, K.M., GOLDSTEIN, R.H., AND WALTON, A.W., 1997, Regional and local controls
ROBERTS, J.W., BARNES, J.J., AND WACKER, H.J., 1976, Subsurface Paleozoic
of diagenesis driven by basin-wide flow system: Pennsylvanian sandstones and
stratigraphy of the northeastern New Mexico basin and arch complex, in Guidebook
limestones, Cherokee Basin, Southeastern Kansas, in Montañez, I., Gregg, J., and
of Vermejo Park, Northeastern New Mexico: New Mexico Geological Society, 27,
Shelton, K., eds., Basin-Wide Diagenetic Patterns: Integrated Petrologic, Geochem-
p. p.141–152.
ical, and Hydrologic Considerations: SEPM, Special Publication 57, p. 235–252.
ROSENBAUM, J., AND SHEPPARD, S.M.F., 1986, An isotopic study of siderites, dolomites
WOODWARD, L.E., 1987, Tectonic framework of northeastern New Mexico and adjacent
and ankerites at high temperatures: Geochimica et Cosmochimica Acta, v. 50,
parts of Colorado, Oklahoma and Texas, in Lucas, S.G., and Hunt, A.P., eds.,
p. p.1147–1150.
Northeastern New Mexico: New Mexico Geological Society, Guidebook 38, p. 67–71.
ROSS, C.A., AND ROSS, J.R.P., 1986, Paleozoic palaeotectonics and sedimentation in WOODWARD, L.E., ANDERSON, O.J., AND LUCAS, S.G., 1999, Late Paleozoic right-slip
Arizona and New Mexico, in Peterson, J.A., ed., Sedimentation in the Rocky faults in the Ancestral Rocky Mountains, in Pazzaglia, F.J., and Lucas, S.G., eds.,
Mountain Region: American Association of Petroleum Geologists, Memoir 41, Albuquerque Geology: New Mexico Geological Society, Guidebook 50, p. 149–154.
p. 653–668. WRIGHT, D.T., 1999, The role of sulfate-reducing bacteria and cyanobacteria in
ROSSI, C., AND CAÑAVERAS, J.C., 1999, Pseudospherulitic fibrous calcite in paleo- dolomite formation in distal ephemeral lakes of the Coorong region, South Australia:
groundwater, unconformity-related diagenetic carbonates (Paleocene of the Ager Sedimentary Geology, v. 126, p. 147–157.
Basin and Miocene of the Madrid Basin, Spain): Journal of Sedimentary Research, WRIGHT, D.T., AND WACEY, D., 2005, Precipitation of dolomite using sulfate-reducing
v. 69, p. 224–238. bacteria from the Coorong Region, South Australia: significance and implications:
SCHRIJVER, K., WILLIAMS-JONES, A., BERTRAND, R., AND CHAGON, A., 1996, Genesis and Sedimentology, v. 52, p. 987–1008.
controls of the Appalachian thrust belt, Quebec, Canada: implications for associated WRIGHT, V.P., AND TUCKER, M.E., 1991, Calcretes: an introduction, in Wright, V.P., and
galena–barite mineralization: Chemical Geology, v. 129, p. 257–279. Tucker, M.E., eds., Calcretes: International Association of Sedimentologists, Reprint
SEARL, A., 1988, Pedogenic dolomites from the oolitic group (Lower Carboniferous) Series, vol. 2: Oxford, U.K., Blackwell, p. 1–22.
South Wales: Geological Journal, v. 23, p. 157–169. WRIGHT, V.P., AND VANSTONE, S.D., 1991, Assessing the carbon dioxide content of
SCOTESE, C.R., 1999, Palaeogeographic Atlas: University of Texas-Arlington, Depart- ancient atmospheres using paleocalcretes: theoretical and empirical constraints:
ment of Geology, PALEOMAP Project Geological Society of London, Journal, v. 148, p. 945–947.
SMITH, L.B, JR, AND DAVIES, G.D., 2006, Structurally controlled hydrothermal alteration WRIGHT, V.P., VANSTONE, S.D., AND MARSHALL, J.D., 1997, Contrasting flooding
of carbonate reservoirs: Introduction: American Association of Petroleum Geologists, histories of Mississippian carbonate platforms revealed by marine alteration effects in
Bulletin, v. 90, p. 1635–1640. palaeosols: Sedimentology, v. 47, p. 825–842.
SMITH, T., AND DOROBEK, S., 1993, Alteration of early-formed dolomite during shallow YOO, C.M., GREGG, J.M., AND SHELTON, K.L., 2000, dolomitization and dolomite
to deep burial: Mississippian Mission Canyon Formation, central to southwestern neomorphism: Trenton and Black river Limestone (Middle Ordovician) northern
Montana: Geological Society of America, Bulletin, v. 105, p. 1389–1399. Indiana, U.S.A.: Journal of Sedimentary Research, v. 70, p. 265–274.
SPÖTL, C., AND WRIGHT, V.P., 1992, Groundwater dolocretes from the Upper Triassic of
the Paris Basin, France: a case study on an arid, continental diagenetic facies:
Sedimentology, v. 39, p. p.1119–1136. Received 18 April 2010; accepted 24 October 2010.

You might also like